< Back to 68k.news TR front page

Near-collapse of the geomagnetic field may have contributed to atmospheric oxygenation and animal radiation in the Ediacaran Period

Original source (on modern site)

Near-collapse of the geomagnetic field may have contributed to atmospheric oxygenation and animal radiation in the Ediacaran Period

Communications Earth & Environment volume 5, Article number: 207 (2024) Cite this article

Abstract

Earth's magnetic field was in a highly unusual state when macroscopic animals of the Ediacara Fauna diversified and thrived. Any connection between these events is tantalizing but unclear. Here, we present single crystal paleointensity data from 2054 and 591 Ma pyroxenites and gabbros that define a dramatic intensity decline, from a strong Proterozoic field like that of today, to an Ediacaran value 30 times weaker. The latter is the weakest time-averaged value known to date and together with other robust paleointensity estimates indicate that Ediacaran ultra-low field strengths lasted for at least 26 million years. This interval of ultra-weak magnetic fields overlaps temporally with atmospheric and oceanic oxygenation inferred from numerous geochemical proxies. This concurrence raises the question of whether enhanced H ion loss in a reduced magnetic field contributed to the oxygenation, ultimately allowing diversification of macroscopic and mobile animals of the Ediacara Fauna.

Similar content being viewed by others

Introduction

The discovery of an ultra-low time-averaged field intensity (UL-TAFI, 0.7 × 1022 A m2) from 565 Ma anorthosites of the Sept Îles Mafic Intrusive Suite (Quebec, Canada) suggested that the geodynamo approached the weak field state where core kinetic energy exceeds magnetic energy1,2. Connections between unusual Ediacaran-Cambrian magnetic fields and evolutionary events3 proposed4 prior to report of the UL-TAFI have been challenged5 (Supplementary Discussion—The time-averaged geomagnetic field). As emphasized by Sagan nearly 60 years earlier6, the atmosphere and ocean can shield life from energetic solar particles. For atmospheric loss controlled by the magnetosphere7 to have played an important role, it must have acted over a relatively long time interval, and/or involved processes amplifying escape, overcoming bounds imposed by the slow upward diffusion of atmospheric constituents8.

The Sept Îles UL-TAFI record is supported by paleointensity investigations of Ediacaran lavas and dikes9. Nevertheless the data have limitations. Thellier analyses of samples dominated by single domain magnetic carriers are the gold standard for paleointensity determination because only these data are solidly based on thermoremanent magnetization theory10. Moreover, because the geomagnetic field exhibits short term secular variation, time-averaging is needed to detect trends reflecting evolving core processes (Supplementary Discussion—The time-averaged geomagnetic field). The Sept Îles Thellier data are time-averaged (~75 kyr) by virtue of the slow cooling of the anorthosites studied1. But this key time-averaged aspect has not yet been replicated with Thellier data from single domain dominated samples of other localities.

Bono et al.1 used single crystal paleointensity (SCP)11 analyses to isolate ideal magnetic records. The feldspar and pyroxene crystals measured contained magnetic needles exsolved along crystallographic planes12. Scanning electron microscopy (SEM) analyses and magnetic hysteresis data confirmed their SD nature. Herein, we examine both the accuracy and duration the UL-TAFI through a comparison of SCP analyses of 2054 Ma (South Africa)13 and 591 Ma (Brazil)14 rocks that come from a layered mafic intrusion and post-collisional magmatic suite, respectively, where prior paleomagnetic and geochronological studies have demonstrated the preservation of primary magnetizations (Supplementary Discussion—Geologic context and prior work). Plagioclase from the rocks studied have single domain magnetic characteristics nearly identical to those of the Sept Îles samples studied by Bono et al.1. Motivated by the results, we then examine anew the potential connections of the geomagnetic field and Ediacaran to Cambrian evolution, and propose linkages with the paleomagnetosphere15,16 and oxygenation17,18,19,20.

Results

We collected eight plagioclase-bearing pyroxenite hand samples in the 2054 ± 1.3 Ma13 Rustenberg Layered Intrusion of the Bushveld Complex in northern South Africa (Supplementary Fig. 1a; "Materials and methods"). Magnetic susceptibility versus temperature curves (K-T; Supplementary Fig. 2a, b; "Materials and methods") indicate that Ti-poor titanomagnetite and pure magnetite are the dominant magnetic carriers in bulk samples (Supplementary Discussion—Rock magnetism of whole rocks). Magnetic hysteresis loops ("Materials and methods") of small bulk-rock specimens (Supplementary Fig. 2c, d) have relatively low coercive force (\(\overline{{B}_{c}}\) = 13.6 ± 8.5 mT), remanent coercivity (\(\overline{{B}_{cr}}\) = 60.8 ± 31.2 mT), and squareness (\(\overline{{M}_{rs}/{M}_{s}}\) = 0.16 ± 0.09; Supplementary Table 1). These data suggest the presence of pseudosingle domain and multidomain magnetic grains (Supplementary Fig. 3a). In contrast, selected optically clear plagioclase grains have wide, open hysteresis loops (Fig. 1a and Supplementary Fig. 2e, f) and high values (\(\overline{{B}_{c}}\) = 69.1 ± 15.7 mT, \(\overline{{B}_{cr}}\) = 107.5 ± 19.7 mT, \(\overline{{M}_{rs}/{M}_{s}}\) = 0.48 ± 0.08) (Supplementary Table 1), indicating the presence of single domain grains (Supplementary Fig. 3a).

Fig. 1: Rock magnetic, SEM, and paleointensity analysis of plagioclase crystals from Bushveld pyroxenites.

a Magnetic hysteresis loop with images of the measured plagioclase crystal (lower right, scale bar 1 mm). b First order reversal curve (FORC) diagram for plagioclase crystal. c, d Backscatter SEM images of elongated magnetic inclusions observed in the plagioclase crystals (inset) and their corresponding EDS spectra in 20 keV. EDS spots analyzed are marked as red circles in the inset images. e, f NRM lost versus TRM gained (circles) and pTRM checks (triangles). All labeled points are °C. Gray circles are steps used in fit (applied field Blab and paleointensity value Banc shown in lower left). Insets: crystal measured shown in top center with 1 mm scale bar. Orthogonal vector plot of field-off steps is shown in upper right. Squares are vertical projection of the magnetization; circles are horizontal projection. Temperature ranges used in the paleointensity fit shown in color.

The extraordinary magnetic recording properties of the plagioclase crystals are further supported by first order reversal curve (FORC) analysis and SEM analyses ("Materials and methods"). FORC diagrams of all clean plagioclase specimens are highly peaked on the Bu = 0 axis and the contours are elongated with negligible vertical spread (Fig. 1b and Supplementary Fig. 4), indicating the presence of noninteracting single domain grains. SEM and energy-dispersive X-ray spectroscopy (EDS) data reveal that iron-titanium oxide needles are omnipresent (Fig. 1c, d). The width of the needles is <500 nm, and the length varies widely, but is often several μm. The presence of Ti further implicates a high temperature of formation12, indicating that the needles can carry a primary thermoremanent magnetization.

We investigated absolute paleointensity by SCP Thellier-Coe analyses utilizing CO2 laser heating21 with partial thermal remanent magnetization (pTRM) checks and selection criteria following prior SCP studies1,22 (Fig. 1e, f and Supplementary Fig. 5, "Materials and methods"). Twenty-two of the 175 crystals measured met selection criteria (17 are category A samples, and 5 are category B samples; Supplementary Table 2). The main cause of rejection was failed pTRM checks. The characteristic remanent magnetization used for paleointensity calculation was isolated between 360 and 580 °C. A single crystal from one hand sample (BQ1-D.14) showed higher unblocking, to 620 °C (Supplementary Fig. 5g and Supplementary Table 2). Categories A and B samples with applied field of 30 μT yield a mean paleointensity of 67.4 ± 3.7 μT (n = 13), and samples with applied field of 60 μT (all category A) define a mean paleointensity of 71.2 ± 4.6 μT (n = 9). These mean values appear to be drawn from the same population (Supplementary Discussion—Applied field bias and recording replication, Supplementary Fig. 6), suggesting no correlation with the strength of the applied TRM23. The average paleointensity from all category A and B samples is 69.0 ± 4.3 μT (n = 22). We correct for cooling rate using a factor of 1.5 (Supplementary Discussion—Paleointensity corrections (cooling rate and anisotropy) which yields a paleointensity of 48.6 ± 3.0 μT (n = 22); we further apply anisotropy corrections ("Materials and methods"; Supplementary Discussion—Paleointensity corrections (cooling rate and anisotropy) leading to a paleointensity estimate of 49.1 ± 2.4 μT (Supplementary Table 3). Using a paleolatitude24 of 45°N (Supplementary Discussion—Geologic context and prior work) a paleomagnetic dipole moment of 8.04 ± 0.40 × 1022 A m2 is calculated.

The Passo da Fabiana Gabbros occur as intrusions in the Dom Feliciano Belt of southern Brazil14. Here we focus on a sample previously studied for geochronology14 ("Materials and methods", Supplementary Fig. 1b, c) for single crystal paleointensity analysis. Sample GPF-01 is a coarse-grained gabbro; U-Pb SHRIMP age data on zircons from sample GPF-01 yield an age14 of 591.2 ± 3.5 Ma. We also investigated another gabbro (sample GPF-110) from the same unit but collected from an outcrop ~1 km away for replication. K-T curves for GPF-01 whole rock samples show the dominant presence of magnetite, along with a FeTi phase (Supplementary Fig. 7a). For GPF-110, these data indicate magnetite, but also a change near 320 °C characteristic of pyrrhotite10 (Supplementary Fig. 7b). Magnetic hysteresis data for groundmass samples of both GPF-01 and GPF-110 show narrow loops reflecting relatively low coercive force, remanent coercivity and squareness (\(\overline{{B}_{c}}\) = 8.2 ± 4.1 mT, \(\overline{{B}_{cr}}\) = 23.1 ± 7.4 mT, \(\overline{{M}_{rs}/{M}_{s}}\) = 0.13 ± 0.08; Supplementary Table 4 and Supplementary Fig. 7c, d), consistent with pseudosingle domain and multidomain carriers, whereas curves for clean feldspars from these rocks define broad open loops (\(\overline{{B}_{c}}\) = 52.8 ± 6.6 mT, \(\overline{{B}_{cr}}\) = 92.3 ± 3.8 mT, \(\overline{{M}_{rs}/{M}_{s}}\) = 0.43 ± 0.06) indicating single domain carriers (Fig. 2a and Supplementary Figs. 3b and 7e, f). FORC diagrams for single crystals show a narrow central ridge (Bu = 0) typical of single domain grains (Fig. 2b and Supplementary Fig. 8). SEM and EDS analyses define the presence of needle-like magnetic inclusions, ≤200 nm in width with variable lengths up to several μm, with varying Fe and Ti content (Fig. 2c, d). No SEM/EDS evidence for pyrrhotite is found in the feldspar samples we analyzed.

Fig. 2: Rock magnetic, SEM, and paleointensity analysis of plagioclase crystals from the Passo da Fabiana Gabbros.

a Magnetic hysteresis loop with images of the measured plagioclase crystal (lower right, scale bar 1 mm). b First order reversal curve (FORC) diagram for plagioclase crystal. c, d Backscatter SEM images of elongated magnetic inclusions observed in the plagioclase crystals (inset) and their corresponding EDS spectra in 20 keV. EDS spots analyzed are marked as red circles in the inset images. e, f NRM lost versus TRM gained (circles) and pTRM checks (triangles). All labeled points are °C. Gray circles are steps used in fit (applied field Blab and paleointensity value Banc shown in lower left). Insets: crystal measured shown in top center with 1 mm scale bar. Orthogonal vector plot of field-off steps is shown in upper right. Squares are vertical projection of the magnetization; circles are horizontal projection. Temperature ranges used in the paleointensity fit are shown in color.

Initial SCP Thellier-Coe experiments with laser heating11,21 indicated remanent fields only a fraction as strong as those preserved by the Bushveld plagioclase crystals, leading us to use lower applied fields (30, 15 and 7.5 μT). Measurement of 29 crystals yielded 21 acceptable results (17 Type A, 4 Type B) with ChRMs and paleointensities isolated between 366 and 548 °C (Fig. 2e, f; Supplementary Fig. 9 and Supplementary Table 5). Most results are from GPF-01, but 3 crystals from GPF-110 give comparable values (Supplementary Discussion—Applied field bias and recording replication, Supplementary Fig. 10). The combined paleointensities obtained are 3.43 ± 0.73, 2.31 ± 0.40 and 2.72 ± 0.60 μT for 30, 15 and 7.5 μT applied fields, respectively. While there is no clear progression in the data, the mean of the value determined using a 30 μT field is higher than that obtained at lower fields, hinting at an applied field bias23. The average paleointensity from all these categories A and B samples using all applied field is 2.73 ± 0.69 μT (n = 21). We use a factor of 1.5 as a cooling rate correction (Supplementary Discussion—Paleointensity corrections (cooling rate and anisotropy), yielding an average paleointensity of 1.82 ± 0.46 μT (n = 21).

Anisotropy measurements and subsequent corrections reveal more clearly that paleointensity values obtained at 30 μT are higher than those obtained using lower applied fields (Supplementary Table 6). Accordingly, we select paleointensity and anisotropy data only from those experiments using applied fields of 15 and 7.5 μT. Applying both the cooling rate correction and anisotropy corrections (Supplementary Table 6) yields a paleointensity of 1.49 ± 0.10 μT. Using a paleolatitude of 42.2°S based on a pole position from the Sierra de las Ánimas complex (Uruguay)25 (Supplementary Discussion—Geologic context and prior work) the paleomagnetic dipole moment is 0.25 ± 0.02 × 1022 A m2.

Discussion

The plagioclase studied from the Bushveld pyroxenite and Passo da Fabiana Gabbros have ideal single domain magnetic behavior yet yield vastly different field strengths. The Bushveld value extends the Archean period of high time-averaged field intensity into the Paleoproterozoic (Fig. 3). In stark contrast, the Passo da Fabiana value is 30 times weaker, and the lowest time-averaged field strength known to date (Fig. 3a, b). When combined with the Sept Îles 565 Ma paleomagnetic dipole moment1, the data suggests the UL-TAFI (defined as fields equal to or less than 10% of the present-day) spanned at least 26 million years. If the UL-TAFI is the nadir of the proposed much longer-term decline of field strength preceding the onset of inner core nucleation1,22 (Supplementary Discussion—Long-term paleointensity history and ICN, Fig. 3), the duration of low (defined here as equal to or less than 20% of the present-day) to ultra-low fields could have been >200 m.y.

Fig. 3: Paleointensity, oxygen and animal evolution.

a Field strength constrained from select Thellier (thermal) SCP studies (blue squares, hexagons) and bulk rock studies (gray squares) updated from Zhou et al.22, with new time-averaged SCP results (red hexagons) reported here. Large squares are time-averaged paleomagnetic dipole moments; small squares are virtual dipole moments (VDMs). Gray circles are select Phanerozoic VDMs from Bono et al.1. Field evolution model (3450 Ma to 565 Ma, red line) is weighted second-order polynomial regression of Precambrian field strength data from Bono et al.1; 565 to 532 Ma trend from Zhou et al.22. b Cryogenian to Cambrian field strength evolution corresponding to dashed rectangle in (a). Open circles are results from non-Thellier methods (non-thermal and thermal) and their sizes are weighted by the numbers of cooling units from Zhou et al.22. Key: green, microwave method; purple, Shaw method; black, Wilson method. Brown open circles are Thellier thermal results. Ultra-low time averaged field interval (UL-TAFI) highlighted by light purple rectangle. Also shown are selenium isotopic data (open symbols) and oxygenation interpretation from Pogge von Strandmann et al.20 (shown here with a 25 my window mean and 1σ error), summary animal radiation of bilaterian and non-bilaterians from Zhuravlev and Wood65), Wood et al.3, Darroch et al.66 and Muscente et al.67, and Shuram excursion ages from Rooney et al.42.

The new data confirming and extending the UL-TAFI strengthen a potential linkage with the Ediacaran evolution of macroscopic animals. While phylogenomic data indicate that the animal kingdom and many phyla may have diverged prior to the Ediacaran Period26,27, paleontological data show that the diversification of macroscopic animals, as archived in the Avalon and White Sea assemblages of the Ediacara Fauna, began at ~575 Ma and reached a climax at ~565 Ma28,29,30,31,32, in a striking temporal correlation with the UL-TAFI (Fig. 3b). Prior explanations for the appearance of the Ediacara Fauna and the Cambrian radiation have centered on genetic, ecological, or environmental driving factors33. The temporal coincidence between the UL-TAFI and the rise of the Ediacara Fauna, however, invites a refocus on environmental factors, particularly atmospheric and oceanic oxygenation events.

Oxygen has long been identified as a key "environmental gatekeeper"34, allowing for evolutionary innovation and for meeting the energy demands of animals35. Although sponges and microscopic animals can survive at low levels of dissolved oxygen (as low as ~1.5 μM, equivalent to 0.5% present atmospheric level [PAL]36,37), macroscopic, morphologically complex, and mobile animals require a greater amount of oxygen to support their metabolic demands38. A complex animal ecosystem involving long food chains and predators requires still greater amounts of oxygen, as indicated by the exclusion of such complex ecosystems from the modern oxygen minimum zone39, which is characterized by dissolved O2 levels of <20-30 μM or <10% PAL. In this regard, it is important to note that mobile animals and macroscopic animals up to decimeters in size have been recorded in the ~565 Ma White Sea assemblage of the Ediacara biota40,41.

Multiple lines of geochemical evidence point to a possible increase in atmospheric and oceanic O2 levels in the late Ediacaran Period, specifically in the period of ~575-565 Ma characterized by negative δ13C values of the Shuram excursion42 (Fig. 3b). These include elevated δ88Mo and δ82/76Se values, as well as [Mo], [U], and [V] concentrations in organic-rich shales deposited approximately at 565 ± 5 Ma and thus coincident with the diversification of macroscopic animals in the White Sea assemblage of the Ediacara biota20,43. Importantly, δ238U data of carbonate rocks, as well as Δ\({}^{{\prime} 17}\)O-δ34S-δ18O of carbonate-associated sulfate, point to multimillion years of oceanic oxygenation that temporally coincide with the Shuram excursion and the Avalon-White Sea assemblages44,45,46. We note that there is much uncertainty and controversy with the regard of the magnitude of this oxygenation event47,48, partly because different geochemical proxies have different sensitivities and hence sometimes give apparently conflicting inferences about oxygen levels. Such uncertainty is further complicated by the dynamic nature of oceanic redox history and spatial heterogeneity of oceanic redox conditions in the late Ediacaran Period49,50,51. However, the concurrence of these independent indicators of oxygenation, paleontological records documenting the rise of the Ediacara Fauna, and paleomagnetic data defining the UL-TAFI merits further exploration for potential linkages.

Hydrogen (H) can escape to space by thermal and non-thermal mechanisms resulting in net oxygenation of the atmosphere52. Changes in climate and/or methane release recorded by Ediacaran isotopic records, including the Shuram excursion42,53 (Fig. 3b), could have led to conditions favoring H loss, but these interpretations are debated54 (Supplementary Discussion—Oxygenation, Ediacaran climate and methane). Here we focus on the potential influence of the UL-TAFI on H loss through changes in the paleomagnetosphere.

Using solar wind evolution models based on solar analogs15, steady-state magnetopause stand-off distances would have been <4.5 Earth radii (r⊕) at 565 Ma1 and <4.2r⊕ at 591 Ma, compared to the present-day value of ~10.7r⊕. During extreme coronal mass ejection events, stand-off distances could have been as low as 1.6r⊕. These smaller standoff distances can lead to more hydrogen escape because they increase the polar cap area of open field lines16 where non-thermal H+ loss occurs today. The steady-state polar cap area of open field lines would have been 3.5 times greater at 591 Ma, and even greater during coronal mass ejection events impacting Earth (Supplementary Discussion—Magnetopause standoff, hydrogen loss and hydrogen supply). Moreover, at the reduced stand-off distances implied by our data, the magnetopause will approach, and even fall below Earth's plasmapause16,55, below which plasma density, dominated by H+, markedly increases (Supplementary Discussion—Magnetopause standoff, hydrogen loss and hydrogen supply). In this case, increased global loss of H+ is expected. Greater global non-thermal H loss is also predicted for the much smaller magnetosphere resulting from the UL-TAFI because the solar wind has greater access to a higher density of H atoms (Supplementary Discussion—Magnetopause standoff, hydrogen loss and hydrogen supply).

Quantifying H loss for Earth with an ultra-low field is challenging as it is observationally inaccessible. The few available models predict greater H loss56, but the predicted change vary depending on assumptions (~30% to 10 times). The highest estimates could result in a few percent change of oxygen (PAL) that might represent a perturbation or crossing of a threshold, allowing Ediacaran animal diversification (Supplementary Discussion—Magnetopause standoff, hydrogen loss and hydrogen supply), but it should be noted that none of these nascent models consider all of the processes discussed above.

An effect that also warrants further consideration is that with decreased magnetopause standoff, greater penetration of highly energetic protons is expected over a larger solid angle, resulting in the formation of NOx compounds which can create ozone holes57. With the greater UV flux (Supplementary Discussion—Magnetopause standoff, hydrogen loss and hydrogen supply), dissociation of water vapor58 is expected, increasing the supply of H moving upward. Thus, the UL-TAFI could have led to both an increased supply of H available for loss, and an increased outflow.

Enhanced H loss from the magnetosphere could have acted over tens or hundreds of millions of years when the geomagnetic field was at low to ultra-low strengths (Fig. 3). The multiple pathways discussed above provide a framework for future studies quantifying hydrogen loss, oxygenation of the atmosphere and the potential consequences on the evolution of complex animals highlighted by the synchroneity of these phenomena and the near collapse of the geodynamo.

Materials and methods

Samples were collected from the Rustenberg Layered Intrusion of the Bushveld Complex, South Africa, at 25° 33.812' S, 27° 32.480' E and from the Passo da Fabiana Gabbros at 31° 38.391' S, 53° 22.172' W (GPF-01) and 31° 37.893' S, 53° 21.815' W (GPF-110). The Rustenberg samples were collected as unoriented hand samples from a quarry exposure ensuring the study of fresh material, whereas the Passo da Fabiana samples are unoriented splits of larger samples used in prior petrological, geochemical and geochronological studies14 ensuring both the study of fresh material and an exact correspondence of age constraints. Rock preparation, rock magnetic and paleointensity analyses were done in the paleomagnetic laboratories at the University of Rochester, and supplemented with select magnetic susceptibility measurements of Passo da Fabiana samples at Michigan Tech. Unoriented single plagioclase crystals were picked by hand from rock crushes. Grains selected for magnetic hysteresis and paleointensity measurements are free of visible inclusions when viewed under ×10 magnification. We used a Geofyzika KLY-4C KappaBridge for the bulk-rock magnetic susceptibility measurements in air at the University of Rochester and a AGICO MFK1-FA (in air) at Michigan Tech. A Princeton Measurements Corporation Alternating Gradient Force Magnetometer was used for magnetic hysteresis measurements. First order reversal curves data59 were processed by FORCinel and VARIFORC software60,61. SEM observation and energy-dispersive X-ray spectroscopy analyses used a Zeiss Auriga Scanning Electron Microscope, in the Integrated Nanosystems Center at the University of Rochester.

Plagioclase grains selected for paleointensity analyses have NRM moments greater than 1 × 10−10 A m2. For the Thellier-Coe paleointensity experiments, samples were mounted on quartz holders having moments ≤1 × 10−12 A m2, and measured using a 2G Enterprises 755 3-component DC SQUID magnetometer with high resolution sensing coils housed in a magnetically shielded room with ambient field <200 nT at the University of Rochester. The non-magnetic purity of mounting material (sodium silicate) has been confirmed by SQUID microscope analyses62. Samples were heated using Synrad v20 and v40 CO2 lasers calibrated using optical pyrometers63.

Paleointensity selection criteria are as follows: Category A samples, (1) the natural remanent magnetization (NRM) versus TRM slope must include at least four steps; (2) the R2 of the NRM/TRM slope fit must be greater than 0.9; (3) the field-off NRM values must trend to the origin of the orthogonal vector plots with a maximum angular dispersion ≤10°; and (4) at least three temperature steps of pTRM checks must be included and the values must be within 15%. Category B samples meet Category A criteria with only one criterion relaxed (Supplementary Tables 2 and 5).

Measurement of TRM anisotropy64 follow the procedures described in Bono et al.1. The sample was first heated to an intermediate unblocking temperature without an applied field. Next, the sample was heated to the same temperature step in a 7.5, 15, 30 or 60 μT magnetic field along six orthogonal axes (+Z, −Z, +X, −X, +Y, and −Y). The magnetization of the sample was measured after each field-on heating. A multidomain tail check was performed after the six field-on steps to test for alteration.

Data availability

Magnetic datasets generated during and/or analyzed in this study are available in the MagIC repository, http://earthref.org/MagIC/19610. Original data published in this manuscript is available at https://doi.org/10.6084/m9.figshare.25343248.

References

  1. Bono, R. K., Tarduno, J. A., Nimmo, F. & Cottrell, R. D. Young inner core inferred from Ediacaran ultra-low geomagnetic field intensity. Nat. Geosci. 12, 143-147 (2019).

    Article  CAS  Google Scholar 

  2. Driscoll, P. E. Simulating 2 Ga of geodynamo history. Geophys. Res. Lett. 43, 5680-5687 (2016).

    Article  Google Scholar 

  3. Wood, R. et al. Integrated records of environmental change and evolution challenge the Cambrian Explosion. Nat. Ecol. Evol. 3, 528-538 (2019).

    Article  Google Scholar 

  4. Meert, J. G., Levashova, N. M., Bazhenov, M. L. & Landing, E. Rapid changes of magnetic field polarity in the late Ediacaran: linking the Cambrian evolutionary radiation and increased UV-B radiation. Gondwana Res. 34, 149-157 (2016).

    Article  Google Scholar 

  5. Lingam, M. Revisiting the biological ramifications of variations in Earth's magnetic field. Astrophys. J. Lett. 874, L28 (2019).

    Article  CAS  Google Scholar 

  6. Sagan, C. Is the early evolution of life related to the development of the Earth's core? Nature 206, 448-448 (1965).

    Article  Google Scholar 

  7. Blackman, E. G. & Tarduno, J. A. Mass, energy, and momentum capture from stellar winds by magnetized and unmagnetized planets: implications for atmospheric erosion and habitability. Mon. Not. R. Astron. Soc. 481, 5146-5155 (2018).

    Article  CAS  Google Scholar 

  8. Hunten, D. Escape of light gases from planetary atmospheres. J. Atmos. Sci. 30, 1481-1494 (1973).

    Article  CAS  Google Scholar 

  9. Shcherbakova, V. et al. Ultra-low palaeointensities from East European Craton, Ukraine support a globally anomalous palaeomagnetic field in the Ediacaran. Geophys. J. Int. 220, 1928-1946 (2020).

    Article  CAS  Google Scholar 

  10. Dunlop, D. J. & Özdemir, Ö. Rock Magnetism: Fundamentals and Frontiers, Vol. 3 (Cambridge University Press, 1997).

  11. Tarduno, J. A., Cottrell, R. D. & Smirnov, A. V. The paleomagnetism of single silicate crystals: recording geomagnetic field strength during mixed polarity intervals, superchrons, and inner core growth. Rev. Geophys. 44, RG1002 (2006).

    Article  Google Scholar 

  12. Feinberg, J. M., Scott, G. R., Renne, P. R. & Wenk, H.-R. Exsolved magnetite inclusions in silicates: features determining their remanence behavior. Geology 33, 513-516 (2005).

    Article  Google Scholar 

  13. Scoates, J. S. & Friedman, R. M. Precise age of the platiniferous Merensky Reef, Bushveld Complex, South Africa, by the U-Pb zircon chemical abrasion ID-TIMS technique. Econ. Geol. 103, 465-471 (2008).

    Article  CAS  Google Scholar 

  14. Dal Olmo-Barbosa, L., Koester, E., Vieira, D. T., Porcher, C. C. & Cedeño, D. G. Crystallization ages of the basic intrusive Ediacaran magmatism in the southeastern Dom Feliciano Belt, southernmost Brazil: implications in the belt geodynamic evolution. J. South Am. Earth Sci. 108, 103143 (2021).

    Article  CAS  Google Scholar 

  15. Tarduno, J. A. et al. Geodynamo, solar wind, and magnetopause 3.4 to 3.45 billion years ago. Science 327, 1238-1240 (2010).

    Article  CAS  Google Scholar 

  16. Siscoe, G. & Chen, C. Paleomagnetosphere. J. Geophys. Res. Space Phys. 80, 4675-4680 (1975).

    Article  Google Scholar 

  17. Canfield, D. E., Poulton, S. W. & Narbonne, G. M. Late-Neoproterozoic deep-ocean oxygenation and the rise of animal life. Science 315, 92-95 (2007).

    Article  CAS  Google Scholar 

  18. Scott, C. et al. Tracing the stepwise oxygenation of the Proterozoic ocean. Nature 452, 456-459 (2008).

    Article  CAS  Google Scholar 

  19. Partin, C. A. et al. Large-scale fluctuations in Precambrian atmospheric and oceanic oxygen levels from the record of U in shales. Earth Planet. Sci. Lett. 369-370, 284-293 (2013).

    Article  CAS  Google Scholar 

  20. Pogge von Strandmann, P. A. E. et al. Selenium isotope evidence for progressive oxidation of the Neoproterozoic biosphere. Nat. Commun. 6, 10157 (2015).

    Article  CAS  Google Scholar 

  21. Tarduno, J. A., Cottrell, R. D., Watkeys, M. K. & Bauch, D. Geomagnetic field strength 3.2 billion years ago recorded by single silicate crystals. Nature 446, 657-660 (2007).

    Article  CAS  Google Scholar 

  22. Zhou, T. et al. Early Cambrian renewal of the geodynamo and the origin of inner core structure. Nat. Commun. 13, 4161 (2022).

    Article  CAS  Google Scholar 

  23. Selkin, P. A., Gee, J. S. & Tauxe, L. Nonlinear thermoremanence acquisition and implications for paleointensity data. Earth Planet. Sci. Lett. 256, 81-89 (2007).

    Article  CAS  Google Scholar 

  24. Letts, S., Torsvik, T. H., Webb, S. J. & Ashwal, L. D. Palaeomagnetism of the 2054 Ma Bushveld Complex (South Africa): implications for emplacement and cooling. Geophys. J. Int. 179, 850-872 (2009).

    Article  Google Scholar 

  25. Rapalini, A. E. et al. The late Neoproterozoic Sierra de las Animas Magmatic Complex and Playa Hermosa Formation, southern Uruguay, revisited: paleogeographic implications of new paleomagnetic and precise geochronologic data. Precambrian Res. 259, 143-155 (2015).

    Article  CAS  Google Scholar 

  26. Erwin, D. H. et al. The Cambrian conundrum: early divergence and later ecological success in the early history of animals. Science 334, 1091-1097 (2011).

    Article  CAS  Google Scholar 

  27. Dohrmann, M. & Woerheide, G. Dating early animal evolution using phylogenomic data. Sci. Rep. 7, 3599 (2017).

    Article  Google Scholar 

  28. Narbonne, G. M. The Ediacara biota: neoproterozoic origin of animals and their ecosystems. Annu. Rev. Earth Planet. Sci. 33, 421-442 (2005).

    Article  CAS  Google Scholar 

  29. Shen, B., Dong, L., Xiao, S. & Kowalewski, M. The Avalon explosion: evolution of Ediacara morphospace. Science 319, 81-84 (2008).

    Article  CAS  Google Scholar 

  30. Xiao, S. & Laflamme, M. On the eve of animal radiation: phylogeny, ecology and evolution of the Ediacara biota. Trends Ecol. Evol. 24, 31-40 (2009).

    Article  Google Scholar 

  31. Droser, M. L., Tarhan, L. G. & Gehling, J. G. The rise of animals in a changing environment: global ecological innovation in the Late Ediacaran. Annu. Rev. Earth Planet. Sci. 45, 593-617 (2017).

    Article  CAS  Google Scholar 

  32. Matthews, J. J. et al. A chronostratigraphic framework for the rise of the Ediacaran macrobiota: new constraints from mistaken point ecological reserve, Newfoundland. GSA Bull. 133, 612-624 (2020).

    Article  Google Scholar 

  33. Knoll, A. H. & Carroll, S. B. Early animal evolution: emerging views from comparative biology and geology. Science 284, 2129-2137 (1999).

    Article  CAS  Google Scholar 

  34. Cloud Jr, P. E. Atmospheric and hydrospheric evolution on the primitive earth: both secular accretion and biological and geochemical processes have affected Earth's volatile envelope. Science 160, 729-736 (1968).

    Article  CAS  Google Scholar 

  35. Payne, J. L. et al. The evolutionary consequences of oxygenic photosynthesis: a body size perspective. Photosynth. Res. 107, 37-57 (2011).

    Article  CAS  Google Scholar 

  36. Sperling, E. A., Halverson, G. P., Knoll, A. H., Macdonald, F. A. & Johnston, D. T. A basin redox transect at the dawn of animal life. Earth Planet. Sci. Lett. 371-372, 143-155 (2013).

    Article  Google Scholar 

  37. Mills, D. B. et al. Oxygen requirements of the earliest animals. Proc. Natl Acad. Sci. 111, 4168-4172 (2014).

    Article  CAS  Google Scholar 

  38. Xiao, S. 6.10—Oxygen and early animal evolution. in Treatise on Geochemistry 2nd edn (eds Holland, H. D. & Turekian, K. K.) 231-250 (Elsevier, 2014).

  39. Sperling, E. A. et al. Oxygen, ecology, and the Cambrian radiation of animals. Proc. Natl Acad. Sci. 110, 13446-13451 (2013).

    Article  CAS  Google Scholar 

  40. Evans, S. D., Gehling, J. G. & Droser, M. L. Slime travelers: early evidence of animal mobility and feeding in an organic mat world. Geobiology 17, 490-509 (2019).

    Article  Google Scholar 

  41. Ivantsov, A., Nagovitsyn, A. & Zakrevskaya, M. Traces of locomotion of Ediacaran macroorganisms. Geosciences 9, 395 (2019).

    Article  Google Scholar 

  42. Rooney, A. D. et al. Calibrating the coevolution of Ediacaran life and environment. Proc. Natl Acad. Sci. 117, 16824-16830 (2020).

    Article  CAS  Google Scholar 

  43. Evans, S. D., Diamond, C. W., Droser, M. L. & Lyons, T. W. Dynamic oxygen and coupled biological and ecological innovation during the second wave of the Ediacara Biota. Emerg. Top. Life Sci. 2, 223-233 (2018).

    Article  CAS  Google Scholar 

  44. Zhang, F. et al. Global marine redox changes drove the rise and fall of the Ediacara biota. Geobiology 17, 594-610 (2019).

    Article  CAS  Google Scholar 

  45. Li, Z. et al. Transient and stepwise ocean oxygenation during the late Ediacaran Shuram Excursion: insights from carbonate δ238U of northwestern Mexico. Precambrian Res. 344, 105741 (2020).

    Article  CAS  Google Scholar 

  46. Wang, H. et al. Sulfate triple-oxygen-isotope evidence confirming oceanic oxygenation 570 million years ago. Nat. Commun. 14, 4315 (2023).

    Article  CAS  Google Scholar 

  47. Mills, D. B. & Canfield, D. E. Oxygen and animal evolution: did a rise of atmospheric oxygen "trigger" the origin of animals? BioEssays 36, 1145-1155 (2014).

    Article  CAS  Google Scholar 

  48. Sperling, E. A. et al. Statistical analysis of iron geochemical data suggests limited late Proterozoic oxygenation. Nature 523, 451-454 (2015).

    Article  CAS  Google Scholar 

  49. Zhang, F. et al. Extensive marine anoxia during the terminal Ediacaran Period. Sci. Adv. 4, eaan8983 (2018).

    Article  Google Scholar 

  50. Evans, S. D. et al. Environmental drivers of the first major animal extinction across the Ediacaran White Sea-Nama transition. Proc. Natl Acad. Sci. 119, e2207475119 (2022).

    Article  CAS  Google Scholar 

  51. Wood, R. A. et al. Dynamic redox conditions control late Ediacaran metazoan ecosystems in the Nama Group, Namibia. Precambrian Res. 261, 252-271 (2015).

    Article  CAS  Google Scholar 

  52. Urey, H. On the early chemical history of the Earth and the origin of life. Proc. Natl Acad. Sci. USA 38, 351-363 (1952).

    Article  CAS  Google Scholar 

  53. Bjerrum, C. J. & Canfield, D. E. Towards a quantitative understanding of the late Neoproterozoic carbon cycle. Proc. Natl Acad. Sci. USA 108, 5542-5547 (2011).

    Article  CAS  Google Scholar 

  54. Busch, J. F. et al. Global and local drivers of the Ediacaran Shuram carbon isotope excursion. Earth Planet. Sci. Lett. 579, 117368 (2022).

    Article  CAS  Google Scholar 

  55. Yamauchi, M. Terrestrial ion escape and relevant circulation in space. Ann. Geophys. 37, 1197-1222 (2019).

    Article  CAS  Google Scholar 

  56. Gunell, H. et al. Why an intrinsic magnetic field does not protect a planet against atmospheric escape. Astron. Astrophys. 614, L3 (2018).

    Article  Google Scholar 

  57. Jackman, C. H., Fleming, E. L., Vitt, F. M. & Considine, D. B. The influence of solar proton events on the ozone layer. Adv. Space Res. 24, 625-630 (1999).

    Article  CAS  Google Scholar 

  58. Nicolet, M. On the photodissociation of water vapour in the mesosphere. Planet. Space Sci. 32, 871-880 (1984).

    Article  CAS  Google Scholar 

  59. Roberts, A., Pike, C. R. & Verosub, K. L. Forc diagrams: a new tool for characterizing the magnetic properties of natural samples. J. Geophys. Res. 105, 461 (2000).

  60. Harrison, R. J. & Feinberg, J. M. FORCinel: an improved algorithm for calculating first-order reversal curve distributions using locally weighted regression smoothing. Geochem. Geophys. Geosyst. 9, 11 (2008).

  61. Egli, R. VARIFORC: an optimized protocol for calculating non-regular first-order reversal curve (FORC) diagrams. Glob. Planet. Change 110, 302-320 (2013).

    Article  Google Scholar 

  62. Tarduno, J. A. et al. Paleomagnetism indicates that primary magnetite in zircon records a strong Hadean geodynamo. Proc. Natl Acad. Sci. 117, 2309-2318 (2020).

    Article  CAS  Google Scholar 

  63. O'Brien, T. et al. Arrival and magnetization of carbonaceous chondrites in the asteroid belt before 4562 million years ago. Commun. Earth Environ. 1, 54-54 (2020).

    Article  Google Scholar 

  64. Veitch, R. J. An investigation of the intensity of the geomagnetic field during Roman times using magnetically anisotropic bricks and tiles. Arch. Sci. Geneve. 37, 359-373 (1984).

    Google Scholar 

  65. Zhuravlev, A. Y. & Wood, R. A. The two phases of the Cambrian Explosion. Sci. Rep. 8, 16656 (2018).

    Article  Google Scholar 

  66. Darroch, S. A. F., Smith, E. F., Laflamme, M. & Erwin, D. H. Ediacaran Extinction and Cambrian explosion. Trends Ecol. Evol. 33, 653-663 (2018).

    Article  Google Scholar 

  67. Muscente, A. D., Boag, T. H., Bykova, N. & Schiffbauer, J. D. Environmental disturbance, resource availability, and biologic turnover at the dawn of animal life. Earth Sci. Rev. 177, 248-264 (2018).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by National Science Foundation (NSF) Grants EAR1656348, 2051550 and 1828866 to J.A.T. We greatly appreciate discussions with Sara Pruss, and acknowledge help from G. Kloc in specimen preparation and B.L. McIntyre in electron microscopy analyses.

Author information

Authors and Affiliations

  1. Department of Earth and Environmental Sciences, University of Rochester, Rochester, NY, 14627, USA

    Wentao Huang, John A. Tarduno, Tinghong Zhou & Rory D. Cottrell

  2. State Key Laboratory of Tibetan Plateau Earth System, Environment and Resources (TPESER), Institute of Tibetan Plateau Research, Chinese Academy of Sciences, 100101, Beijing, China

    Wentao Huang

  3. Department of Physics & Astronomy, University of Rochester, Rochester, NY, 14627, USA

    John A. Tarduno & Eric G. Blackman

  4. Laboratory for Laser Energetics, University of Rochester, Rochester, NY, 14623, USA

    John A. Tarduno & Eric G. Blackman

  5. Geological Sciences, University of KwaZulu-Natal, Durban, 4000, South Africa

    John A. Tarduno & Michael K. Watkeys

  6. Department of Geosciences, University of Arizona, Tucson, AZ, 85721, USA

    Mauricio Ibañez-Mejia

  7. Instituto de Geociências, Universidade Federal do Rio Grande do Sul, Porto Alegre, RS, Brazil

    Laércio Dal Olmo-Barbosa & Edinei Koester

  8. Department of Geological and Mining Engineering and Sciences, Michigan Technological University, Houghton, MI, 49931, USA

    Aleksey V. Smirnov & Gabriel Ahrendt

  9. Department of Physics, Michigan Technological University, Houghton, MI, 49931, USA

    Aleksey V. Smirnov

  10. Department of Earth and Environmental Sciences, Lehigh University, Bethlehem, PA, 18015, USA

    Kenneth P. Kodama

  11. Department of Earth, Ocean and Atmospheric Science, Florida State University, Tallahassee, FL, 32306, USA

    Richard K. Bono

  12. Geomagnetism Laboratory, University of Liverpool, Liverpool, L69 3GP, UK

    Richard K. Bono

  13. Code 670, NASA/Goddard Space Flight Center, Greenbelt, MD, 20771, USA

    David G. Sibeck

  14. State Key Laboratory for Mineral Deposits Research, Institute of Continental Geodynamics, School of Earth Science and Engineering, Nanjing University, Nanjing, 210046, China

    Yong-Xiang Li

  15. Department of Earth and Planetary Sciences, University of California, Santa Cruz, CA, 95064, USA

    Francis Nimmo

  16. Department of Geosciences, Virginia Tech, Blacksburg, VA, 24061, USA

    Shuhai Xiao

Contributions

W.H. collected paleointensity, rock magnetic and electron microscope data, analyzed the results, contributed to implications for the dynamo, ICN and animal radiations, and wrote the manuscript with contributions from all authors. J.A.T. conceived the study as part of a larger project with A.V.S., K.P.K. and M.I.M., conducted field work in South Africa, analyzed paleointensity data and rock magnetic data, contributed consideration of paleointensity history and ICN, magnetic shielding, atmospheric loss and potential linkages to animal radiation, and wrote the manuscript with contributions from all the authors. T.Z. contributed electron microscope analyses and to considerations of paleointensity history. M.I.M. contributed to the geochronologic and geological interpretations and overall project design. L.D.B. conducted field studies in South America, collected samples and provided geologic context including geochronology. E.K. provided geologic context including geochronology for samples from South America. E.G.B. contributed theoretical consideration on magnetic shielding and atmospheric loss. A.V.S. contributed magnetic susceptibility data and to overall project development. G.A. contributed magnetic susceptibility data. R.D.C. analyzed paleointensity and rock magnetic data. K.P.K. contributed to the overall project development, paleointensity history, and ICN. R.K.B. contributed to paleointensity history, field stability, ICN and data analysis. Y.-X.L. contributed to paleointensity history, field stability and ICN. F.N. contributed to the consideration of implications for the dynamo and ICN. S.X. provided paleobiology/paleoenvironment context. M.K.W. conducted field work in South African and contributed geologic interpretations.

Corresponding author

Correspondence to John A. Tarduno.

Ethics declarations

Competing interests

The authors declare no competing interests. Field work and sample collection was conducted with or by individuals associated with institutions in South Africa or Brazil with appropriate permissions for geological sampling. Specifically, field work in South Africa was conducted by J.A.T. and M.K.W. Field work in South America was conducted by L.D.B.

Peer review

Peer review information

Communications Earth & Environment thanks Joseph Meert and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editor: Carolina Ortiz Guerrero. A peer review file is available.

Additional information

Publisher's note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Cite this article

Huang, W., Tarduno, J.A., Zhou, T. et al. Near-collapse of the geomagnetic field may have contributed to atmospheric oxygenation and animal radiation in the Ediacaran Period. Commun Earth Environ 5, 207 (2024). https://doi.org/10.1038/s43247-024-01360-4

Download citation

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

< Back to 68k.news TR front page